Aminoxyl-Catalyzed Electrochemical Diazidation of Alkenes Mediated

Jan 23, 2019 - We report the development of a new aminoxyl radical catalyst, CHAMPO, for the electrochemical diazidation of alkenes. Mediated by an ...
0 downloads 0 Views 533KB Size
Subscriber access provided by MIDWESTERN UNIVERSITY

Communication

Aminoxyl-Catalyzed Electrochemical Diazidation of Alkenes Mediated by a Metastable Charge-Transfer Complex Juno C. Siu, Joseph B. Parry, and Song Lin J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 23 Jan 2019 Downloaded from http://pubs.acs.org on January 24, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Aminoxyl-Catalyzed Electrochemical Diazidation of Alkenes Mediated by a Metastable ChargeTransfer Complex Juno C. Siu‡, Joseph B. Parry‡, Song Lin* Department of Chemistry and Chemical Biology, Cornell University, Ithaca, New York 14853, United States Supporting Information Placeholder

ABSTRACT: We report the development of a new aminoxyl radical catalyst, CHAMPO, for the electrochemical diazidation of alkenes. Mediated by an anodically generated charge-transfer complex in the form of CHAMPO–N3, radical diazidation was achieved across a broad scope of alkenes without the need for a transition metal catalyst or a chemical oxidant. Mechanistic data support a dual catalytic role for the aminoxyl serving as both a single-electron oxidant and a radical group transfer agent.

The discovery of reactions mediated by organic radicals continues to provide solutions to challenging synthetic problems in traditional two-electron chemistry.1 In this context, design and implementation of new catalytic strategies have both expanded the toolbox available for accessing new synthetic targets and transformed the fundamental understanding of reactions involving openshell pathways.2 For example, persistent aminoxyl radicals [e.g., (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO)] have been extensively explored in catalytic oxidation reactions with both conventional chemical3 and electrochemical techniques,4 which has given rise to synthetically useful processes for small-molecule and polymer syntheses. Despite significant advances, we contend that the scope of TEMPO chemistry remains to be fully explored. TEMPO and related N-oxyl radicals can undergo one-electron redox processes, granting access to three discrete oxidation states.4 This feature distinguishes these radicals from common organic compounds and likens them to many transition metal complexes. In this fashion, TEMPO has been shown to enable single-electron oxidation events in an inner-sphere manner via the formation of metastable closed-shell intermediates.5,6 Nevertheless, the systematic use of the “metallic” character of N-oxyls in catalyst development remains meager.7 To date, reactions catalyzed by N-oxyls are largely confined to oxidations of alcohols,8 aldehydes,9 amines,10 (thio)amides,11 and peroxyl radicals.12 In these transformations, the persistent radical is used primarily in two capacities—as a single/two-electron oxidant or an H atom abstractor. In this report, we expand the scope of N-

oxyl catalysis in the development of metal-free electrochemical diazidation of alkenes. We previously reported a first-generation electrochemical protocol for the diazidation of alkenes catalyzed by Mn (Scheme 1A, Eq 1).13 This work was built on foundational contributions by Minisci,14 Magnus,15 Snider,16 Xu,17 and others18 in alkene diazidation using conventional chemical methods. These reactions provide 1,2-diazides in a single step from common alkenes, giving access to vicinally dinitrogenated structures that are highly prevalent in biomedically and synthetically relevant molecules.19 Nevertheless, all methods reported to date require the use of a transition metal or a chemical oxidant, or both. These reactive species often lead to environmentally deleterious wastes and present an explosion hazard when used alongside N3–. Although our electrochemical protocol13 circumvents the use of strong chemical oxidants, the need for a metal catalyst and a Brønsted acid (HOAc) as the sacrificial oxidant remains undesirable from a safety perspective. Scheme 1. Mechanistic hypothesis for N-oxyl-catalyzed alkene diazidation (A) Background: Electrochemical diazidation and azidooxygenation MnBr2 (cat.) Diazidation

NaN3 R

via

electrolysis

N3 R

N3 (Eq 1)

[MnIII]–N3

TEMPO Azidooxygenation via TEMPO–N3

OTMP N3 (Eq 2)

R TMP = 2,2,6,6tetramethylpiperidyl

For seminal developments in chemical diazidation, see refs. 14–18. For seminal developments in chemical azidooxygenation, see ref. 20b.

(B) Working mechanistic hypothesis for aminoxyl (NO•)-catalyzed alkene diazidation e–

NO•

N3–

via:

ref. 20a

N N

N3•

R R

N3 (I)

N

N O

TEMPO–N3 CTC

TEMPO–N3

R

OTMP N3 R Azidooxygenation (undesired; TEMPO is consumed)

TEMPO

–TEMPO N3

N3

N3 R Diazidation (desired; catalytic in TEMPO)

While studying the electrochemical alkene diazidation, we discovered a new electrochemical method for the

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

azidooxygenation of alkenes mediated by TEMPO (Scheme 1A, Eq 2).20 Detailed mechanistic investigation showed that TEMPO plays two roles in this reaction. Aside from being a radical trap, TEMPO also promotes the single-electron oxidation of N3– to N3●.21 This process is mediated by a metastable charge-transfer complex (CTC) formed between anodically generated TEMPO+ and N3–. Upon azidyl addition, the nascent carbon-centered radical I rapidly combines with TEMPO to deliver the masked vicinal aminoalcohol (Scheme 1B, black-red pathway). We envisioned that the TEMPO-consuming azidooxygenation reaction could be engineered into a TEMPO-catalyzed alkene diazidation if intermediate I was terminated by an N3● equivalent instead of directly by TEMPO (Scheme 1B, black-blue pathway). Specifically, we hypothesize that the TEMPO–N3 adduct may behave as an azidyl group transfer agent akin to the function of a redoxactive transition metal complex (e.g., [MnIII]–N3). This operation will furnish the desired vicinal diazide while returning the aminoxyl from an oxoammonium ion to a neutral radical oxidation state. We investigated the plausibility of N-oxyl-catalyzed electrochemical diazidation (Table 1). In our alkene azidooxygenation, diazidation products were occasionally observed in small quantities when sterically encumbered alkenes were used. The diazide formation was more pronounced at a higher applied voltage and using less TEMPO. Under these conditions, the [TEMPO+]/[TEMPO] ratio is maximized to favor the formation of the CTC and mitigate the undesired azidooxygenation. For example, the electrolysis of alkene 1a, NaN3, and 1 equiv TEMPO in MeCN/H2O at a cell voltage of 2.7 V22 yielded 36% 1,2diazide 2a in addition to 45% 3a (entry 1). Decreasing TEMPO loading to a catalytic amount still provided encouragingly 11% 2a alongside 7% 3a (entry 2). However, the catalyst was unproductively consumed in the azidooxygenation pathway, which was also indicated by the color change of the reaction medium. Upon starting electrolysis, the solution turned dark red, which is characteristic of the TEMPO–N3 CTC. This color disappeared after a mere 10 minutes as a result of TEMPO depletion. This visual indicator thus provided a convenient readout for reaction optimization. Table 1. Reaction optimizationa

We reasoned that heating the reaction would promote homolysis of the C–O bond in byproduct 3a, thus returning both TEMPO and the carbon-centered radical to the catalytic cycle.23 Counterintuitively, this operation led to rapid decomposition of the CTC and suppressed diazidation. Instead, cooling the reaction improved the yield to 46% at 0 °C and 61% at –10 °C (entries 3, 4). However, the color of the CTC decayed after 2 h with concomitant formation of 3a (ca. 5%). Studies by Bobbitt,8k Minteer/Sigman24, and Stahl25 showed that the distal C4 substituent of the aminoxyl has a profound influence on its reactivity in alcohol oxidation. Accordingly, we surveyed various catalysts26 and found that 4-acetamido-TEMPO (ACT) proved the most promising, furnishing 71% 2a and 99% Me

N3 9

N3

Me N3

Me N3 2r, 77% (dr = 1:1 at C9)c

Me Me N3

Ph

O 2

Reactive functional groups:

N3 Me

Me N3

Me

N3

MeO

N3 2q, 82% (dr = 1:1)

O

O NHFmoc 2l, 81% (dr = 1:1)

Me N3

2n, 55% (dr = 5:1)

Br

N3

2j, 82%

Me Me

N3

Ph

N3

2m, 79%

N3

O

2k, >99%

N3

Me

N3

2i, 72%, dr = 11:1

Me N3

N3

Me

2f, 85%

Ph 2h, 88%

2a, 97%

BocN

N3

N

2g, 79%b

TBSO

N3

2e, 68% N3

N3

MeO

Me N3

O N3

N

OMe

2d, 49%b

N3 Me

2

PhO

8

2c, 70%b

N3

R

LiClO4 (0.1 M), H2O/MeCN (1:12) C(+)/Pt(–), Ecell = 2.7 V, –10°C

1

O

N3

N3 Fe Me N3

N3

2u, 60%d

2v, 97%

O N

N3 Me N 2w, 84%

Me

N3

N3

Me

Me N3

H

2x, 90% (dr = 1:1)

N3

N3

Me

Ts N 2y, 70%

Product derivatization: NH3OTs Fe Me NH3OTs 4, 72%e

Me N3 Ph 5, 52%f

NHAc

TBSO

Me NHBoc NHBoc 6, 63%g

aConditions:

1 (0.2 mmol, 1 equiv), CHAMPO (10 mol%), NaN3 (3 equiv), LiClO4 (1.8 equiv), MeCN/H2O (3.8 mL, 12:1). b15 mol% CHAMPO. cLimonene-1,2-oxide (1r) was purchased as a pair of diastereomers at C1,2. dDetermined by 1H NMR owing to the volatility of the product. eConditions: 2v, PPh3, H2O/THF, reflux; then pTSA. fConditions: 2k, PPh3, H2O/THF, reflux; then Ac2O, NEt3, DCM, 0 °C. gConditions: 2a, H , Pd/C, EtOAc, Boc O. 2 2

The 1,2-diazide products can be readily reduced to the corresponding vicinal diamines using Staudinger reduction or catalytic hydrogenation. Notably, the different steric environments of the two azido groups in 2k enabled chemoselective reduction to form β-azidoamide 5. The diazides could also be readily converted to bistriazoles or 1,2-dinitro compounds,36 which are common structures in energetic materials. Compared with our first-generation Mn-catalyzed alkene diazidation, the new protocol presents several practical advances. For example, replacing the transition metal catalyst with an organocatalyst and the protic acid (HOAc) with H2O eliminates the generation of hazardous metal azides and hydrazoic acid. These features make the diazide synthesis safer and more amenable to practical applications.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

We conducted further experiments to exemplify the user friendliness of the protocol (Table 3). Using inexpensive Ni foam as the cathode instead of Pt only marginally decreased product yield (entry 2). Using carbon as both cathode and anode did not provide 2k in appreciable amounts. To make the reaction system metal-free, we carried out electrolysis in a divided cell that completely separates the diazidation from the cathodic proton reduction on Pt (entry 3). The electrolysis was also successfully performed using commercial ElectraSyn 2.037 or AA batteries as the power source (entries 4, 5). Electrolyte choice proved unimportant (entry 6). In fact, given the ionic nature of NaN3, electrolysis in the absence of an additional electrolyte yielded 2k quantitatively (entry 7). Finally, diazidation on a synthetically relevant scale resulted in a promising isolated yield of 69% (entry 8). Table 3. Optimization of the electrolysis setupa Me Ph

“Standard conditions”: CHAMPO (10 mol%) NaN3 (1.5 equiv per C–N) LiClO4 (0.1 M), MeCN/H2O (12:1) C(+)/Pt(–), Ecell = 2.7 V, –10 °C

1k Entry

Variation from “standard conditions”

Me N3 Ph

(A) Cationic pathway I: Direct oxidation of carbon-centered radical I N3

R

(I)

CHAMPO+ or anode single-electron oxidation

R

>99

2

Ni(–) instead of Pt(–)

92

3

Divided cell, constant current of 8 mA

81

4

ElectraSyn 2.0 as power source, 0.3 mmol 1k

82

5

AA batteries (x2) as power source

89

6

Using TBAPF6 instead of LiClO4

95

7

No LiClO4, Ecell = 3.4 Vb

>99

8

1 mmol scale

O R

N(R’)2

anode

N3

oxidation–mesolytic 3 cleavage (C) Radical pathway 2H2O

R

69c

conditions: see Table 1, entry 5. higher cell voltage is applied to compensate for the increased solution resistance between cathode and anode. cIsolated yield; reaction was carried out at a higher substrate concentration (see SI);

Aside from the radical pathway in our initial working mechanistic hypothesis (see Scheme 1B), two additional reaction pathways may be plausible to account for diazidation reactivity. Carbon-centered radical I may undergo single-electron oxidation, either directly on the anode or by CHAMPO+, to generate the corresponding carbocation (II; Scheme 2A). The cation would subsequently react with N3– to form the diazide. Alternatively, azidooxygenated product 3 could be an intermediate en route to the diazide via oxidation and mesolytic cleavage38 to form II (Scheme 2B). Scheme 2. Plausible mechanistic pathways

N3

N3–

N3

N3

R

(II)

2e– N3–

H2 + 2OH– cathode

substrate CHAMPO+ –e–

CHAMPO–N3

Innersphere SET cycle

R

slow N3•

CHAMPO

N3

bA

N3

R

(II)

(B) Cationic pathway II: Oxidation of the azidooxygenation product 3

CHAMPO –e–

Yield (%)

None

N3

N3–

N3

2k

1

aReaction

Page 4 of 8

N3 (I)

fast Group transfer cycle

CHAMPO+

anode

R

CHAMPO–N3

N3 R

N3

product

N3–

Experimental evidence (Scheme 3) is consistent with our original mechanistic proposal in which the aminoxyl–N3 adduct mediates the radical diazidation, whereas for certain types of substrates, carbocation pathway I also might operate. Reactions with radical clocks showed that cyclopropane 7 and bisallylamine 9/10 underwent radicalinduced ring rupture or 5-exo-trig cyclization, respectively. In all cases, no direct alkene diazidation was observed. In principle, upon radical ring opening or cyclization, the resultant carbon-centered radical could be further oxidized to the cation before nucleophilic azidation; however, these carbocations, either highly electron-deficient or at a primary carbon, would be rather unstable. We also carried out diazidation of amide 13, which—if operating via a carbocation mechanism—would cyclize to furnish oxazoline 15.39 However, electrolysis produced diazide 14 cleanly as the only observable product. Together, these data strongly support the radical mechanism. Scheme 3. Mechanistic probes. Structures in purple are expected products if the second C–N3 formation were via a cationic pathway.

ACS Paragon Plus Environment

Page 5 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society accessing N3● from N3–.21 In light of our new mechanistic information, we conclude that the CHAMPO–N3 CTC plays two key roles—promoting the generation of free N3● and mediating the second azidation through a putative radical group transfer event. Preliminary data suggest that the CTC is the resting state of the catalyst under a sufficient anodic potential. The first C–N3 bond formation is slower than the second owing to the relatively slow, endergonic fragmentation of the CTC to CHAMPO and N3●. This rate difference allows the two distinct radical addition events to take place successively with high chemoselectivity toward diazidation over side pathways that unproductively consumes I. In this mechanistic hypothesis, the metal-like behavior of N-oxyl radicals in the alkene diazidation is particularly intriguing. Our ongoing efforts will be devoted to further evaluating the mechanistic proposal and using the knowledge to guide the discovery of other synthetically useful transformations.

ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website, including experimental procedures and characterization data (PDF) and crystallographic data for CHAMPO (CCDC1881714) (CIF). We further studied alkene substrates that would lead to resonance-stabilized carbocation intermediates (Scheme 3C) and observed radical-to-carbocation oxidation. For example, camphor-derived alkene 16 was converted to 11% diazide 17 in addition to 42% 1,3-azidoalcohol 18 via carbocation rearrangement. We reasoned that the norbornyl structure selectively stabilizes carbocation (II) but not its radical precursor (I), thus favoring the oxidation of I to II. Notably, neither rearrangement-azidation (19) or direct 1,2azidohydroxylation products were observed. This result strongly supports the hypothesis that the alkene diazidation undergoes predominantly a radical pathway,17b,40 whereas the azidohydroxylation adopts a cationic pathway that triggers the rearrangement.41 Furthermore, 4-methoxy-αmethylstyrene (1g) was converted to diazide 2g (major) and azidoalcohol 20 (minor). A key finding from these experiments is that if a carbocation is formed, its capture by H2O is inevitable. Importantly, the majority of the substrates we investigated (see Table 2) produced no appreciable amounts of azidoalcohol products. This piece of information lends further support for the radical mechanism involving Noxyl-mediated azidyl transfer. Finally, we conducted a control experiment using azidooxygenated product 21 instead of the alkene as the substrate for electrolysis and observed no diazide formation. This result excluded the possibility that N-oxyl adducts are key intermediates in the alkene diazidation (see Scheme 2B). Accordingly, we propose a CHAMPO–N3-mediated electrocatalytic cycle (Scheme 2C). We previously showed that N-oxyl radicals trigger an inner-sphere oxidation pathway that substantially lowers the potential required for

AUTHOR INFORMATION Corresponding Author

*[email protected] Author Contributions ‡These

authors contributed equally.

Notes

The authors declare no competing financial interests.

ACKNOWLEDGMENT Financial support was provided by Cornell University and NIGMS (R01GM130928). S.L. thanks the ONR for a Young Investigator Award which supports the synthesis of energetic materials. This study made use of the NMR facility supported by the NSF (CHE-1531632) and the ESR facility supported by NIGMS (P41GM103521). We thank IKA for the donation of ElectraSyn 2.0, Dr. Samantha MacMillan for help with X-ray crystallography analysis, Drs. Ivan Keresztes and Frank Schroeder for help with NMR spectral analysis, Yifan Shen for reproducing experimental results, and Dean Kim for experimental assistance.

REFERENCES 1. (a) Renaud, P.; Sibi, M. P. Radicals in Organic Synthesis; Wiley-VCH: Weinhein, 2001. (b) Yan, M.; Lo, J. C.; Edwards, J. T.; Baran, P. S. Radicals: Reactive Intermediates with Translational Potential. J. Am. Chem. Soc. 2016, 138, 12692. (c) Ischay, M. A.; Yoon,

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

T. P. Accessing the Synthetic Chemistry of Radical Ions. Eur. J. Org. Chem. 2012, 2012, 3359. (d) Moeller, K. M. Using Physical Organic Chemistry To Shape the Course of Electrochemical Reactions. Chem. Rev. 2018, 118, 4817. 2. (a) Studer, A.; Curran, D. P. Catalysis of Radical Reactions: A Radical Chemistry Perspective. Angew. Chem., Int. Ed. 2016, 55, 58. (b) Skubi, K. L.; Blum, T. R.; Yoon, T. P. Dual Catalysis Strategies in Photochemical Synthesis. Chem. Rev. 2016, 116, 10035. (c) Narayanam, J. M. R.; Stephenson, C. R. Visible light photoredox catalysis: applications in organic synthesis. J. Chem. Soc. Rev. 2011, 40, 102. (d) Francke, R.; Little, R. D. Redox catalysis in organic electrosynthesis: basic principles and recent developments. Chem. Soc. Rev. 2014, 43, 2492. (e) Yan, M.; Kawamata, Y.; Baran, P. S. Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance. Chem. Rev. 2017, 117, 13230. (f) Crossley, S. W. M.; Martinez, R. M.; Obradors, C.; Shenvi, R. A. Mn-, Fe-, and CoCatalyzed Radical Hydrofunctionalizations of Olefins. Chem. Rev. 2016, 116, 8912. 3. (a) Bobbitt, J. M.; Brückner, C.; Merbouh, N. Oxoammonium and Nitroxide-Catalyzed Oxidation of Alcohols. Org. React. 2009, 74, 103. (b) Wertz, S.; Studer, A. Nitroxide-catalyzed transition-metal-free aerobic oxidation processes. Green Chem. 2013, 15, 3116. (c) Leadbeater, N. E.; Bobbitt, J. M. TEMPO-Derived Oxoammonium Salts as Versatile Oxidizing Agents. Aldrichim. Acta 2014, 47, 65. 4. Nutting, J. E.; Rafiee, M. R.; Stahl, S. S. Tetramethylpiperidine N-Oxyl (TEMPO), Phthalimide N-Oxyl (PINO), and Related N-Oxyl Species: Electrochemical Properties and Their Use in Electrocatalytic Reactions. Chem. Rev. 2018, 118, 4834. 5. (a) Qian, X.-Y.; Li, S.-Q.; Song, J.; Xu, H.-C. TEMPOCatalyzed Electrochemical C–H Thiolation: Synthesis of Benzothiazoles and Thiazolopyridines from Thioamides. ACS Catal. 2017, 7, 2730. (b) Peterson, B. M.; Lin, S.; Fors, B. P. Electrochemically Controlled Cationic Polymerization of Vinyl Ethers. J. Am. Chem. Soc. 2018, 140, 2076. (c) Furukawa, K.; Shibuya, M.; Yamamoto, Y. Chemoselective Catalytic Oxidation of 1,2-Diols to αHydroxy Acids Controlled by TEMPO–ClO2 Charge-Transfer Complex. Org. Lett. 2015, 17, 2282. 6. For examples of stable TEMPO–X-type adducts formed via single-electron transfer, see: (a) Li, Y.; Pouliot, M.; Vogler, T.; Renaud, P.; Studer, A. α-Aminoxylation of Ketones and β-Chloro-αaminoxylation of Enones with TEMPO and Chlorocatecholborane. Org. Lett. 2012, 14, 4474. (b) Wang, X.; Ye, Y.; Zhang, S.; Feng, J.; Xu, Y.; Zhang, Y.; Wang, J. Copper-Catalyzed C(sp3)–C(sp3) Bond Formation Using a Hypervalent Iodine Reagent: An Efficient Allylic Trifluoromethylation. J. Am. Chem. Soc. 2011, 133, 16410. 7. TEMPO has been used (stoichiometrically or catalytically) in lieu of transition metals for formal C–C cross coupling. See: (a) Murarka, S.; Studer, A. Transition Metal‐Free TEMPO‐Catalyzed Oxidative Cross‐ Coupling of Nitrones with Alkynyl‐Grignard Reagents. Adv. Synth. Catal. 2011, 353, 2708. (b) Maji, M. S.; Murarka, S.; Studer, A. Transition-Metal-Free Sonogashira-Type Coupling of ortho-Substituted Aryl and Alkynyl Grignard Reagents by Using 2,2,6,6-Tetramethylpiperidine-N-oxyl Radical as an Oxidant. Org. Lett. 2010, 12, 3878. 8. For examples, see: (a) Semmelhack, M. F.; Chou, C. S.; Cortes, D. A. Nitroxyl-mediated electrooxidation of alcohols to aldehydes and ketones. J. Am. Chem. Soc. 1983, 105, 4492. (b) Hoover, J.; Stahl, S. Highly Practical Copper(I)/TEMPO Catalyst System for Chemoselective Aerobic Oxidation of Primary Alcohols. J. Am. Chem. Soc. 2011, 133, 16901. (c) Badalyan, A.; Stahl, S. S. Cooperative electrocatalytic alcohol oxidation with electron-proton-transfer mediators. Nature 2016, 535, 406–410. (d) Liu, R.; Liang, X.; Dong, C.; Hu, X. Transition-Metal-Free:  A Highly Efficient Catalytic Aerobic Alcohol Oxidation Process. J. Am. Chem. Soc. 2004, 126, 4112. (e) Shibuya, M.; Osada, Y.; Sasano, Y.; Tomizawa, M.; Iwabuchi, Y. Highly Efficient, Organocatalytic Aerobic Alcohol Oxidation. J. Am. Chem. Soc. 2011, 133, 6497. (f) Shibuya, M.; Doi, R.; Shibuta, T.; Uesugi, S.; Iwabuchi, Y. Organocatalytic One-Pot Oxidative Cleavage of Terminal Diols to Dehomologated Carboxylic

Acids. Org. Lett. 2012, 14, 5006. (g) Hickey, D.; McCammant, M.; Giroud, F.; Sigman, M.; Minteer, S. Hybrid Enzymatic and Organic Electrocatalytic Cascade for the Complete Oxidation of Glycerol. J. Am. Chem. Soc. 2014, 136, 15917. (h) Zhao, M.; Li, J.; Mano, E.; Song, Z.; Tschaen, D.; Grabowski, E.; Reider, P. Oxidation of Primary Alcohols to Carboxylic Acids with Sodium Chlorite Catalyzed by TEMPO and Bleach. J. Org. Chem. 1999, 64, 2564. See also ref. 5c. Stoichiometric oxidation: (i) Bailey, W. F.; Bobbitt, J. M. Mechanism of the Oxidation of Alcohols by Oxoammonium Cations. J. Org. Chem. 2007, 72, 4504. (j) Ganem, B. Biological spin labels as organic reagents. Oxidation of alcohols to carbonyl compounds using nitroxyls. J. Org. Chem. 1975, 40, 1998. (k) Bobbitt, J. M. Oxoammonium Salts. 4-Acetylamino-2,2,6,6-tetramethylpiperidine-1-oxoammonium Perchlorate:  A Stable and Convenient Reagent for the Oxidation of Alcohols. Silica Gel Catalysis. J. Org. Chem. 1998, 63, 9367. 9. For examples, see: (a) Miles, K.; Abrams, M.; Landis, C.; Stahl, S. KetoABNO/NOx Cocatalytic Aerobic Oxidation of Aldehydes to Carboxylic Acids and Access to α -Chiral Carboxylic Acids via Sequential Asymmetric Hydroformylation/Oxidation. Org. Lett. 2016, 18, 3590. (b) Rafiee, M.; Konz, Z.; Graaf, M.; Koolman, H.; Stahl, S. Electrochemical Oxidation of Alcohols and Aldehydes to Carboxylic Acids Catalyzed by 4-Acetamido-TEMPO: An Alternative to “Anelli” and “Pinnick” Oxidations. ACS Catal. 2018, 8, 6738. 10. For a review, see: (a) Bartelson, A.; Lambert, K.; Bobbitt, J.; Bailey, W. Recent Developments in the Nitroxide‐Catalyzed Oxidation of Amines: Preparation of Imines and Nitriles. ChemCatChem. 2016, 8, 3421-3430. For examples, see: (b) Lennox, A.; Goes, S.; Webster, M.; Koolman, H.; Djuric, S.; Stahl, S. Electrochemical Aminoxyl-Mediated α-Cyanation of Secondary Piperidines for Pharmaceutical Building Block Diversification. J. Am. Chem. Soc. 2018, 140, 11227. (c) Semmelhack, M. F.; Schmid, C. R. Nitroxyl-mediated electro-oxidation of amines to nitriles and carbonyl compounds. J. Am. Chem. Soc. 1983, 105, 6732. (d) Kim, J.; Stahl, S. Cu/Nitroxyl-Catalyzed Aerobic Oxidation of Primary Amines into Nitriles at Room Temperature. ACS Catal. 2013, 3, 1652. (e) Sonobe, T.; Oisaki, K.; Kanai, M. Catalytic aerobic production of imines en route to mild, green, and concise derivatizations of amines. Chem. Sci. 2012, 3, 3249. (f) Wu, Y.; Yi, H.; Lei, A. Electrochemical Acceptorless Dehydrogenation of N-Heterocycles Utilizing TEMPO as OrganoElectrocatalyst. ACS Catal. 2018, 8, 1192. (g) Romero, N. A.; Margrey, K. A.; Tay, N. E.; Nicewicz, D. A. Site-selective arene C-H amination via photoredox catalysis. Science 2015, 349, 1326. (h) Somfai, P.; Yu, L. Synthesis of α-Keto Amides by a Pyrrolidine/TEMPO-Mediated Oxidation of α-Keto Amines. Synlett 2016, 27, 2587. (g) Wertz, S.; Studer, A. Metal‐Free 2,2,6,6‐Tetramethylpiperidin‐1‐yloxy Radical (TEMPO) Catalyzed Aerobic Oxidation of Hydroxylamines and Alkoxyamines to Oximes and Oxime Ethers. Helv. Chim. Acta 2012, 95, 1758. 11. (a) Tarantino, K. T.; Miller, D. C.; Callon, T. A.; Knowles, R. R. Bond-Weakening Catalysis: Conjugate Aminations Enabled by the Soft Homolysis of Strong N–H Bonds. J. Am. Chem. Soc. 2015, 137, 6440. See also refs. 5a, 5b. Stoichiometric reaction: (b) Xu, F.; Zhu, L.; Zhu, S.; Yan, X.; Xu, H.-C. Electrochemical Intramolecular Aminooxygenation of Unactivated Alkenes. Chem. Eur. J. 2014, 20, 12740. 12. Griesser, M.;Shah, R.; Van Kessel, T. A.; Zilka, O.; Haidasz, A. E.; Pratt, A. D. The Catalytic Reaction of Nitroxides with Peroxyl Radicals and Its Relevance to Their Cytoprotective Properties. J. Am. Chem. Soc. 2018, 140, 3798. 13. Fu, N.; Sauer, G.; Saha, A.; Loo, A.; Lin, S. Metal-catalyzed electrochemical diazidation of alkenes. Science 2017, 357, 575. 14. Minisci, F. Free-radical additions to olefins in the presence of redox systems. Acc. Chem. Res. 1975, 8, 165. 15. (a) Magnus, P., Roe, M. B. and Hulme, C. New trialkylsilyl enol ether chemistry: direct 1,2-bis-azidonation of triisopropylsilyl enol ethers: an azido-radical addition process promoted by TEMPO. J. Chem. Soc., Chem. Commun. 1995, 263. (b) Magnus, P.; Lacour, J.; Evans, P. A.; Roe, M. B.; Hulme, C. Hypervalent Iodine Chemistry:  New Oxidation Reactions Using the Iodosylbenzene−Trimethylsilyl

ACS Paragon Plus Environment

Page 6 of 8

Page 7 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Azide Reagent Combination. Direct α- and β-Azido Functionalization of Triisopropylsilyl Enol Ethers. J. Am. Chem. Soc. 1996, 118, 3406. 16. Snider, B. B.; Lin, H. An Improved Procedure for the Conversion of Alkenes and Glycals to 1,2-Diazides Using Mn(OAc)3·2H2O in Acetonitrile Containing Trifluoroacetic Acid. Synth. Commun. 1998, 28, 1913. 17. (a) Yuan, Y. A.; Lu, D. F.; Chen, Y. R.; Xu, H. Iron-Catalyzed Direct Diazidation for a Broad Range of Olefins. Angew. Chem. Int. Ed. 2016, 55, 534. (b) Shen, S.-J.; Zhu, C.-L.; Lu, D.-F.; Xu, H. IronCatalyzed Direct Olefin Diazidation via Peroxyester Activation Promoted by Nitrogen-Based Ligands. ACS Catal. 2018, 8, 4473. 18. For examples, see: (a) Fumagalli, G.; Rabet, P. T. G.; Boyd, S.; Greaney, M. F. Three‐Component Azidation of Styrene‐Type Double Bonds: Light‐Switchable Behavior of a Copper Photoredox Catalyst. Angew. Chem. Int. Ed. 2015, 54, 11481. (b) Lu, M.-Z.; Wang, C.-Q.; Loh, T.-P. Copper-Catalyzed Vicinal Oxyazidation and Diazidation of Styrenes under Mild Conditions: Access to Alkyl Azides. Org. Lett. 2015, 17, 6110. (c) Peng, H.; Yuan, Z.; Chen, P.; Liu, G. Palladium‐Catalyzed Intermolecular Oxidative Diazidation of Alkenes. Chin. J. Chem. 2017, 35, 876. (d) Fristad, W.E; Brandvold, T. A.; Peterson, J. R.; Thompson, S. R. Conversion of alkenes to 1,2-diazides and 1,2-diamines. J. Org. Chem. 1985, 50, 3647. (e) Moriarty, R. M.; Khosrowshahi, J. S. A versatile synthesis of vicinal diazides using hypervalent iodine. Tetrahedron Lett. 1986, 27, 2809. Related azidoamination: (f) Shen, K.; Wang, Q. Copper-Catalyzed Alkene Aminoazidation as a Rapid Entry to 1,2-Diamines and Installation of an Azide Reporter onto Azahetereocycles. J. Am. Chem. Soc. 2017, 139, 13110. 19. (a) Lucet, D.; Le Gall, T.; Mioskowski, C. The Chemistry of Vicinal Diamines. Angew. Chem. Int. Ed. 1998, 37, 2580. (b) Chemler, S. R.; Fuller, P. H. Heterocycle synthesis by copper facilitated addition of heteroatoms to alkenes, alkynes and arenes. Chem. Soc. Rev. 2007, 36, 1153. 20. (a) Siu, J. C.; Sauer, G. S.; Saha, A.; Macey, R. L.; Fu, N.; Chauvirie, T.; Lancaster, K. L.; Lin, S. Electrochemical Azidooxygenation of Alkenes Mediated by a TEMPO–N3 ChargeTransfer Complex. J. Am. Chem. Soc. 2018, 40, 12511. For related seminal work on alkene azidooxygenation using a non-electrochemical method, see: (b) Zhang, B.; Studer, A. Stereoselective Radical Azidooxygenation of Alkenes. Org. Lett. 2013, 15, 4548. 21. The potential for azidyl formation is lowered by ca. 200 mV. See ref 20a. 22. This corresponds to ca. 0.7 V anodic potential vs. Fc/Fc+ (ferrocene/ferrocenium). 23. Studer, A.; Harms, K.; Knoop, C.; Müller, C.; Schulte, T. New Sterically Hindered Nitroxides for the Living Free Radical Polymerization:  X-ray Structure of an α-H-Bearing Nitroxide. Macromolecules 2004, 37, 27. 24. Hickey, D.; Schiedler, D.; Matanovic, I.; Doan, P.; Atanassov, P.; Minteer, S.; Sigman, M. Predicting Electrocatalytic Properties: Modeling Structure–Activity Relationships of Nitroxyl Radicals. J. Am. Chem. Soc. 2015, 137, 16179. 25. (a) Gerken, J.; Pang, Y.; Lauber, M.; Stahl, S. Structural Effects on the pH-Dependent Redox Properties of Organic Nitroxyls: Pourbaix Diagrams for TEMPO, ABNO, and Three TEMPO Analogs. J. Org. Chem. 2018, 83, 7323. (b) Rafiee, M; Miles, K. C.; Stahl, S. S. Electrocatalytic Alcohol Oxidation with TEMPO and Bicyclic Nitroxyl Derivatives: Driving Force Trumps Steric Effects. J. Am. Chem. Soc. 2015, 137, 14751. 26. For a full list of catalysts surveyed, see SI.

27. (a) Ma, Z.; Huang, Q.; Bobbitt, J. M. Oxoammonium salts. 5. A new synthesis of hindered piperidines leading to unsymmetrical TEMPO-type nitroxides. Synthesis and enantioselective oxidations with chiral nitroxides and chiral oxoammonium salts. J. Org. Chem. 1993, 58, 4837. (b) Moad, G.; Rizzardo, E.; Solomon, D. H. The reaction of acyl peroxides with 2,2,6,6-tetramethylpiperidinyl-1-oxy. Tetrahedron Lett. 1981, 22, 1165. 28. This strategy has been employed in nitroxide-mediated polymerization (NMP) to reverse N-oxyl trapping and thus decrease the temperature needed to maintain chain growth. See ref. 23. 29. In addition to increased steric encumbrance, CHAMPO also has fewer β-H’s. The cyclohexane motifs could also stabilize the CHAMPO+ ion through hyperconjugation. See ref. 27a. 30. Haugland, M.; El-Sagheer, A.; Porter, R.; Peña, J.; Brown, T.; Anderson, E.; Lovett, J. 2′-Alkynylnucleotides: A Sequence- and Spin Label-Flexible Strategy for EPR Spectroscopy in DNA. J. Am. Chem. Soc. 2016, 138, 9069. See also ref. 26a. CHAMPO’s E1/2 = +0.29 V. 31. When the electrolysis was left continued for 12 h instead, the color remained persisted in the end of the reaction. 32. Schäfer, H. Oxidative Addition of the Azide Ion to Olefins. A Simple Route to Diamines. Angew. Chem. Int. Ed. 1970, 9, 158. 33. For instance, CAN gave